Bioactive Fibers and Polymers - American Chemical Society

for the treatment of chronic dermal ulcers exceeds $3 billion in the ..... Surgery Research Laboratory and the nursing staff of the V C U / M C V Woun...
0 downloads 0 Views 2MB Size
Chapter 5

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

The Non-Healing Wound: Mechanisms and Dressings

Dorne R. Yager, Robert F. Diegelmann, I. Kelman Cohen, and Benedict C. Nwomeh Plastic Surgery Research Laboratory, Virginia Commonwealth University,

Richmond, VA 23298 Chronic dermal ulcers are a major cause of disability and have a significant socioeconomic impact. Annual health care expenditures for the treatment of chronic dermal ulcers exceeds $3 billion in the United States alone. The development of improved treatment strategies should proceed from a better understanding of the pathophysiology of these problem wounds. A consistent feature of chronic leg and pressure ulcers is the presence of large numbers of activated neutrophils and perhaps other inflammatory cells. These cells, especially neutrophils, generate large amounts of degradative enzymes and oxygen metabolites that may overwhelm endogenous controls leading to tissue damage. This article summarizes studies that have focused on characterizing the proteolytic environment of chronic dermal ulcers and the implications that this may have for current and future treatment strategies of chronic nonhealing wounds.

Introduction The skin is the body's largest and most complex organ. It performs numerous functions including acting as a barrier against infection, maintaining fluid balance, and providing thermoregulation. Insults that result in a disruption of the continuity of the skin permit entry of microrganisms and are potentially life threatening. A s a result, we have evolved the ability to make rapid and somewhat less than perfect patches or scar tissue to fill these breaches. This process of skin wound healing or repair consists of an extraordinarily complex cascade of highly regulated biochemical and cellular events. It involves many types of cells that not only interact with one and another but must also respond in a coordinated fashion with the extracellular matrix and soluble factors such as cytokines and growth factors. 64

© 2001 American Chemical Society

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

65 The wound healing cascade can be perceived as four overlapping phases. A n initial or hemostasis phase, an early or inflammatory phase, an intermediate or fibroplasia phase, and the late or remodeling phase. Each phase is characterized by distinct biologic processes. The hemostasis phase consists of vasoconstriction and fibrin clot formation. This serves to stop further hemorrhage and to place a temporary plug in the gap of the skin. This plug serves dual functions in it inhibits the invasion of microorganisms and serves as a provisional scaffold upon which cells involved in the subsequent phases use for a pathway to migrate into the wound site. During the inflammatory stage, phagocytic leukocytes consisting of neutrophils followed by macrophages ingest invading microorganisms, foreign bodies and devitalized tissue. These leukocytes also produce soluble factors that appear to attract and coordinate the activities of the cells involved in the fibroplasia phase. During the fibroplasia phase, reepithelialization, angiogenesis (ingrowth of new blood vessels), and connective tissue deposition occur. By far, collagen is the predominant protein of connective tissues (25% of all body protein and more than 50% of the protein in scar tissue)(l). Although peak levels of collagen occur relatively early during repair this new collagen provides little functional strength to the wound. During the remodeling phase, there is a continual rearrangement and cross-linking of collagen that ultimately leads to a mature scar with a strength approaching but not equaling that found in normal skin. A n important feature of this rearrangement is an ongoing equilibrium of collagen degradation and synthesis. However, wound healing does not always occur in this predictable fashion. Many local and systemic factors possess the capability to influence the pattern and rate of healing. A chronic wound is one that fails to heal in a timely fashion because of the existence of one or more pathologic conditions. These include infection, hypoxia, radiation damage, venous insufficiency, toxins, malnutrition, diabetes, pharmacological agents (e.g., steroids), and age. Chronic dermal ulcers represent a major perplexing and costly clinical problem, a problem that will only worsen as our population grows older. Therefore, there has been a great deal of interest in developing new strategies for treating these problem wounds. A consistent feature of arterial, stasis, and pressure ulcers (which make up the majority of chronic wounds) is a chronic, prolonged inflammatory phase (2,3). A n example of the degree of inflammation seen in many chronic wounds is shown in Figure 1. To facilitate their egress from the circulation, penetrate the extracellular matrix, and degrade and digest devitalized tissue and pathogens, neutrophils possess an armamentarium of proteolytic and free radical generating enzymes (4). Unlike other cells that primarily express and secrete proteases on demand, the proteases of neutrophils are formed during myelopoietic development and stored within a variety of cytoplasmic granules and secretory vesicles (5). Although neutrophils have obvious important positive roles in host defense and debridement of damaged tissues, these cells with their free radical generating enzymes and proteases have been implicated in mediating much of the tissue damage associated with chronic inflammatory diseases such as periodontal disease, rheumatoid arthritis, adult respiratory distress syndrome, and cystic fibrosis (6-12). We have thus proposed that a similar over exuberant neutrophil response may likewise contribute to a significant extent in the pathophysiology of chronic wounds (13,14).

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

66

Figure J. Inflammatory nature of chronic wounds. Representative H&E stai of a pressure ulcer granulation tissue. Large numbers of neutrophils ar distributed throughout the granulation tissue as well as with edematous ( blood vessels. Lumens of vessels (arrow) contain an abundance of migr neutrophils. Also note apparent widespread fragmentation of collagen extracellular matrix. Photograph courtesy ofBrad Garrett.

Results & Discussion Neutrophil Proteases in Chronic Wounds The matrix metalloproteases (matrixins, MMPs) are a family of approximately 20 mammalian neutral p H proteases that collectively, can degrade virtually all the protein components of the extracellular matrix (15). The M M P s are expressed as zymogens that require cleavage of an amino-terminal domain to become active. One subgroup of MMPs, the interstitial collagenases (MMP-1 or collagenase-1, M M P - 8 or collagenase-2, and MMP-13 or collagenase-3), are the only mammalian enzymes capable of initiating the degradation of the triple helix of native fibrillar collagens (including types I, Π, III, VII, and X ) (16). Thus, it is the actions of these enzymes that represent the rate limiting step in the turnover of the major protein component of the extracellular matrix. Because it is made almost exclusively by neutrophils, M M P 8 is often called neutrophil collagenase. MMP-8 is sequestered in the specific or secondary granules of neutrophils. E L I S A analysis reveals that on a molar basis,

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

67

M M P - 8 is by far the most abundant collagenase in both healing wounds and in chronic wounds (Table I) (17). Chronic wound fluids contain significantly higher levels of collagenolytic activity than is found in fluids from surgical wounds, healing open dermal wounds, or split thickness wounds (Table I) (17,18). Based on substrate preference studies, the active collagenase found in chronic wounds is composed almost entirely of MMP-8 (Figure 2). Substrate gel electrophoretic techniques have been used to indicate that levels of the gelatinase, M M P - 9 , are higher in chronic wound fluids compared to fluids from healing wounds (18-21). When performed under quantitative conditions, levels of M M P - 9 have been found to be 25-fold higher in chronic wound fluids than surgical wound fluids (Table I). Although not exclusively expressed by neutrophils, M M P - 9 is a major constituent of the small storage or tertiary granules stromelysins, M M P - 3 and MMP-10, in chronic wounds with most of the expression associated with basal keratinocytes.

Table I. Levels of Proteases and Protease Inhibitors in Wounds Chronic Wound Open Dermal PlasmaSurgical Drainage Fluids Wound Fluids Fluids Collagenolytic Activity" N.D. 462±31 20.6±9.2 MMP-8? N.D. 60±7.5 40115 3.0±3.0 MMP-9" N.D. 1,8631591 72+19 Elastase Activity/ N.D. 1.1011.2 0.087±0.08 TIMP 130125 125+20 5,000±25 MMP/TlMP 51+25 16±10 6.815 a.2-Macroglobulin 53.6±20 23.8±15 f

1

b

h

e

"Mean levels of where one unit cleaves 1 μg of type I collagen in 60 minutes at 30°C. Mean levels in micrograms MMP-8 ±SEM per milligram total protein . Determined by sandwich ELISA. Mean levels in nanograms ±SEM per milligram total protein. Déterminai by substrate gel electrophoresis. Mean levels in milliunits ±SD per milligram total protein. Determined by proteolysis of methoxysuccinyl-ala-ala-pro-val-p-nitroanilide. *Mean levels in micrograms per milligram protein ±SD. Determined by quantitative immunoblotting 6

c

rf

Neutrophils also contain several cationic serine proteases (5). The enzyme, elastase, is particularly abundant The amount of elastase contained in the granules of a single neutrophil is in excess of one picogram (4) Several groups have used specific inhibitors as well as antisera to demonstrate that elastase or an elastase-like protease is abundant in chronic wound fluids (13,22-24). Mean levels of elastase activity is approximately 12-fold higher in fluids of chronic dermal ulcers compared to surgical

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

68

Figure 2. MMP-8 is the active collagenase in chronic wounds. Panel A. MM and MMP-8 exhibit substrate preference in assays using soluble collage substrate. In lanes 1-3, type I ateliocollagen was incubated with 1) buffer 1, or 3) MMP-8. In lanes 4-6, type III ateliocollagen was incubated with 2) MMP-1, or 3) MMP-8. Type I collagen serves as a preferential substrat MMP-8 whereas, type III collagen is the preferred substrate for MMP-1. Pan Two representativefluidsfrom chronic wounds were incubated with type I c (lanes 1-4) or type III collagen (5-8) and the products resolves by electrophoresis. Odd numbered lanes represent woundfluidincubated collagen in the presence of EDTA (specific inhibitor of MMPs). Both chron woundfluidsgenerated TC productsfromtype I collagen substrate (lanes 2 a with little corresponding activity directed towards type III collagen (lanes This is consistent with the majority of the collagenolytic activity in chronic being due to neutrophil (MMP-8) collagenase. 4

wound fluids (Table I). Elastase possesses a broad substrate specificity, preferentially cleaving bonds that are carboxy-terminal to valine and to a lesser extent alanine (25). Because of this, elastase has an extremely broad substrate specificity. This enzyme has been demonstrated to be responsible for degrading a number of extracellular matrix proteins including fibrin, fibronectin, tenascin, and vitronectin (2,22-24,26,27). The extracellular matrix may not be the only significant target for elastase in chronic wounds. As shown in Figure 3, elastase present in chronic wounds is capable of degrading critical soluble factors such as platelet-derived growth factor and transforming growth factor βι.(13,28). Although it has not been examined within the milieu of chronic dermal ulcers, neutrophil elastase has also been shown to degrade cell surface receptor proteins (Tierney and Yager, unpublished results) (29-31). Additional serine proteases have been found in chronic wounds and neutrophils probably are also the primary source of these enzymes. These include cathepsin G , another cationic serine protease with broad substrate specificity, urokinase-type plasminogen activator (uPA), and protease 3 (3,32-35). In all probability, when looked for, elevated levels of additional neutrophil proteases will be found in chronic dermal ulcers.

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

69

Figure 3. Neutrophil elastase in chronic woundfluidsis capable of degrad peptide growth factors. I-Transforming growth factor-^ j was incubated (lane 1), with purified neutrophil elastase (lanes 2 and 3), and with 10 chronic woundfluid(lanes 4 and 5) at 37°C for 90 minutes. The presence intact growth factor was assessed by gel electrophoresis and autoradiogr specific inhibitor of neutrophil elastase (10 μΜ N-methoxysuccinyl-ala-ala val-chloromethyl ketone) was included in lanes 3 and 5. 12S

Antiprotease Screen The tissue-destructive potential of neutrophil-derived enzymes are countered by powerful plasma antiproteinases. These include a proteinase inhibitor, a macroglobulin, leukoproteinase inhibitor, plasminogen activator inhibitor (PAI-1), and tissue inhibitor of metalloproteinases (TIMPs) (4). Together, these inhibitors create an antiprotease shield in both plasma and in interstitial fluids. There is growing evidence that this antiprotease shield can be overwhelmed by the over exuberant neutrophil response in chronic wounds. M M P s are specifically inhibited by TIMPs, a family of small proteins (36). TIMPs inhibit active MMPs by forming tight non-covalent 1:1 complexes (37). E L I S A analysis indicates that TIMP-1 levels are lower in fluids from leg and pressure ulcers than are found at peak levels in fluids from healing surgical wounds (Table I) (17). Similarly, E L I S A analysis has also demonstrated the existence of elevated levels of MMP-1/TIMP-1 complexes in fluids from pressure ulcers (18). a -Macroglobulin, is an abundant (2.5 mg/ml) plasma protein that is a potent nonspecific protease inhibitor (38). In healing wounds, levels of this proteinase inhibitor approach that of plasma (13,22,39). In chronic dermal ulcer fluids, proteolytic activity has been shown to correlate with a decrease in the level of intact a -macroglobulin monomers or with the appearance of the protease cleaved forms (Table I) (13,22,39,40). Complexes of a -macroglobulin and MMP-1 have also been observed in stasis ulcer fluids (40). r

2

2

2

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

2

70

Another abundant plasma protease inhibitor, α ι-antiprotease ( α ι-antitrypsin), irreversibly inhibits neutrophil elastase by forming an enzyme-inhibitor complex (38). Levels of complexed forms of a antiprotease are significantly increased in chronic wound fluids and at least some of these complexes contain elastase (13,22). r

Plasma levels for a -macroglobulin and aj-antiprotease are micromolar and similar levels are found in fluids of healing surgical or open dermal wounds (13,22). In healing wounds, TIMP-1 levels increase more than ten-fold to approximately 1,500 ng/ml (18,19). This raises the question of how this seemingly substantial and extensive antiprotease shield can be overwhelmed in chronic wounds? In all probability, the continual influx of large numbers of neutrophils into the chronic wound site places great strain on the antiprotease shield. A s discussed above, neutrophils contain enormous amounts of proteases. In addition, these proteases may cooperate in a synergistic fashion to reduce the amounts of endogenous protease inhibitors within the wound site. TIMPs are substrates for neutrophil elastase, whereas α ι-antiprotease itself is a substrate for neutrophil collagenase (MMP-8) (41,42). Another important neutrophil function that can influence the proteaseantiprotease balance is the generation of hypochlorous acid (HOC1) and N chloramines via the myeloperoxidase-H202-halide system. HOC1 and N-chloramines efficiently oxidize or chlorinate a wide number of molecules including both a2macroglobulin and α 1-antiprotease (43,44). Up to 2500 to 5000 nmol of HOC1 can be generated by as few as 25 χ 106 neutrophils (45). Furthermore, the ability of H O G and N-chloramines to activate latent MMPs may shift the protease-antiprotease equilibrium further towards degradation (46,47).

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

2

Treatment of Chronic Dermal Ulcers The implications of an over exuberant neutrophil response within the milieu of a wound are various and significant. The net deposition of extracellular matrix is difficult to achieve in face of overwhelming proteolytic environment. Extracellular matrix fragments generated by the action of the neutrophil proteases can act as chemotactic peptides that attract additional waves of neutrophils (48-50). Soluble growth factors and their receptors are both likely to be degraded further reducing the ability of wounds to heals or to respond to the application of exogenous growth factors. Thus, the accumulating evidence indicating the ability of the neutrophil to mediate damage to normal tissue would seem to indicate a need to develop therapeutic interventions that can ameliorate these unwanted destructive activities. Potential strategies could target neutrophil functions, the activity of specific neutrophil products or by augmentation and protection of the antiprotease shield. Anti-inflammatory agents that block or inhibit one or more neutrophil functions such as chemotaxis, adherence, infiltration, or degranulation may be worth exploring (5158). Topical application of a chemically modified tetracycline to full-thickness open wounds in streptozotocin treated rats results in decreased collagenase and gelatinase levels and increased granulation tissue formation (59). Protection of protease inhibitors from oxidative damage is another potential approach. A methionine at

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

71

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

position 358 is the primary target for oxidative inactivation of oti-antiprotease (60). Free methionine protects this site from oxidative damage (61). General free radical scavengers might also be employed to not only protect serine protease inhibitors but to also reduce levels of oxidant-activated MMPs. Likewise, inhibitors of enzymes involved in generating reactive oxygen metabolites might be utilized to modulate the chronic wound environment (61-63). Augmentation of the antiprotease shield could be produced by inducing the expression of endogenous antiproteases or by topical application. Alternatively, natural or synthetic exogenous protease inhibitors could be used (64-66).

R'RNH

Figure 4. Schematic of interactions of neutrophil enzymatic and oxidative pr and endogenous protease inhibitors. Short-lived and long-lived oxidants gen by the neutrophil NADPH oxidase/myeloperoxidase system can activate laten proteases (MMP-8 and MMP-9) as well as inactivate major protease inhibito as TIMP, OLrmacroglobulin, and aj-protease inhibitor. Furthermore, neutroph proteases can also interact in a synergistic fashion as evidenced by the abil MMP-8 to degrade dj-protease inhibitor and elastase to degrade TIMP.

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

72

Summary

A growing body of evidence supports a role for neutrophils in the pathophysiology of chronic dermal ulcers. There is a growing appreciation that the extensive proteolytic and oxidative armamentarium of the neutrophil can subvert the extensive and overlapping barriers that have been erected to protect host tissues from injury. A better understanding of that role and of the functions of neutrophils will provide opportunities for novel treatment strategies of these problem wounds. The development and use of inhibitors directed against neutrophil functions and neutrophil products may be useful in attenuating inflammatory damage. These reagents may be effective by themselves or as adjuncts to other treatment strategies. One obvious potential use would be to alter the environment of a chronic dermal ulcer so as to make therapies involving peptide growth factors more effective. There is a bewildering array of dressing products for treating chronic dermal ulcers on the market. With few exceptions, the functions of these dressings are very limited that basically, do one or more of three things: provide protection, facilitate wound debridement, or regulate wound moisture. Perhaps with our increased understanding of the pathophysiology of chronic wounds the stage has been set for thé next generation of chronic wound dressings. These "smart dressings" would modulate the biochemical environment of chronic wounds add or remove enzymes or oxidants. This might be achieved by relatively nonspecific approaches which could include using resins that act as ion exchange supports or by modifying the dressing materials themselves. Alternatively, therapeutic agents could be incorporated into dressing materials. These could include specific inhibitors of proteases, antioxidants, and agents that interfere with neutrophil functions. There would be many advantages to these approaches. Treatment would be topical rather than systemic. B y varying the composition of the dressing materials, the dynamic nature of wound repair could be taken into account. Acknowledgements. This work was supported in part by National Institutes of Health grant GM20298 and GM-47566. The authors also thank the staff of the Plastic Surgery Research Laboratory and the nursing staff of the V C U / M C V Wound Healing Center. D . R . Y . also thanks the residents of Islay for their unceasing spirituous support.

Literature Cited 1. 2. 3. 4. 5.

Nimni M E . Semin Arthritis Rheum 1974; 4: 95-150. Palolahti M , Lauharanta J, Stephens R W , Kuusela P, Vaheri A . Exp Dermatol 1993; 2: 29-37. Rogers AA, Burnett S, Moore JC, Shakespeare P G , Chen W Y J . Wound Repair Regen 1995; 3: 273-283. Weiss SJ. Ν Engl J Med 1989; 320: 365-376. Borregaard N . Curr Opin Hematol 1996; 3: 11-18.

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

73 6. 7. 8. 9. 10. 11.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34.

Malech H L , Gallin JI. Ν Engl J Med 1987; 317: 687-694. Henson P M , Johnston RBJr. J Clin Invest 1987; 79: 669-674. Anderson B O , Brown J M , Harken A H . J Surg Res 1991; 51: 170-179. Meyer K C , Zimmerman J. J Lab Clin Med 1993; 121: 654-661. Birrer P. Respiration 1995; 62 Suppl 1: 25-28. Garcia JG, James H L , Zinkgraf S, Perlman M B , Keogh B A . A m Rev Respir Dis 1987; 136:811-817. Sandholm L . J Clin Periodontol 1986; 13: 19-26. Yager DR, Chen S M , Ward SI, Olutoye OO, Diegelmann RF, Cohen IK. Wound Repair Regen 1997; 5: 23-32. Nwomeh B C , Yager DR, Cohen IK. Clin Plast Surg 1998; 25: 341-356. Nagase H, Woessner JF. Journal O f Biological Chemistry 1999; 274: 2149121494. Welgus H G , Jeffrey JJ, Eisen A Z . J Biol Chem 1981; 256: 9511-9515. Nwomeh B C , Liang H X , Cohen IK, Yager DR. J Surg Res 1999; 81: 189-195. Yager DR, Zhang LY, Liang H X , Diegelmann RF, Cohen IK. J Invest Dermatol 1996; 107: 743-748. Bullen E C , Longaker M T , Updike D L , Benton R, Ladin D, Hou Z , Howard E W . J Invest Dermatol 1995; 104: 236-240. Wysocki A B , Staiano-Coico L , Grinnell F. J Invest Dermatol 1993; 101: 64-68. Tarlton JF, Bailey A J , Crawford E, Jones D, Moore K , Harding K G . Wound Repair Regen 1999; 7: 347-355. Grinnell F , Zhu M. J Invest Dermatol 1996; 106: 335-341. Rao C N , Ladin D A , Liu Y Y , Chilukuri K , Hou Z Z , Woodley D T . J Invest Dermatol 1995; 105: 572-578. Herrick S, Ashcroft G , Ireland G , Horan M, McCollum C , Ferguson M. Lab Invest 1997; 77: 281-288. Owen C A , Campbell EJ. J Leukoc Biol 1999; 65: 137-150. Berman M, Manseau E , Law M, Aiken D. Invest Ophthalmol Vis Sci 1983; 24: 1358-1366. Latijnhouwers M A , Bergers M, Veenhuis R T , Beekman B , Ankersmit-Ter Horst M F , Schalkwijk J. Arch Dermatol Res 1998; 290: 490-496. Wlaschek M , Peus D, Achterberg V , Meyer-Ingold W , Scharffetter-Kochanek K . Br J Dermatol 1997; 137: 646-646. Porteu F , Brockhaus M , Wallach D , Engelmann H , Nathan C F . J Biol Chem 1991; 266: 18846-18853. Döring G , Frank F, Boudier C , Herbert S, Fleischer B, Bellon G . J Immunol 1995; 154: 4842-4850. Harnacher J, Sadallah S, Sehifferli JA, Villard J, Nicod LP. Eur Respir J 1998; 11: 112-119. Vaalamo M, Weckroth M, Puolakkainen P, Kere J, Saarinen P, Lauharanta J, Saarialho-Kere UK. Br J Dermatol 1996; 135: 52-59. Stacey M C , Burnand KG, Mahmoud-Alexandroni M, Gaffney PJ, Bhogal BS. B r J Surg 1993; 80: 596-599. Wysocki A B , Kusakabe A O , Chang S, Tuan T L . Wound Repair Regen 1999; 7: 154-165.

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

74

35. He Y , Young PK, Grinnell F. J Invest Dermatol 1998; 110: 67-71. 36. Wojtowicz-Praga S M , Dickson R B , Hawkins MJ. Invest New Drugs 1997; 15: 61-75. 37. Willenbrock F, Murphy G . A m J Respir Crit Care Med 1994; 150: S165-S170 38. Travis, J. and Salvesen G.S. Ann Rev Biochem 1983; 52:655-709. 39. Witte M B , Thornton FJ, Kiyama T, Efron D T , Schulz GS, Moldawer LL, Barbul A . Surgery 1998; 124: 464-470. 40. Grinnell F, Zhu M, Parks W C . J Invest Dermatol 1998; 110: 771-776. 41. Itoh Y, Nagase H. J Biol Chem 1995; 270: 16518-16521. 42. Michaelis J, Vissers M C , Winterbourn C C . Biochem J 1990; 270: 809-814. 43. Reddy VY, Desorchers PE, Pizzo SV, Gonias SL, Sahakian JA, Levine R L , Weiss SJ. J Biol Chem 1994; 269: 4683-4691. 44. Weiss SJ, Regiani S. J Clin Invest 1984; 73: 1297-1303. 45. Wasil M, Halliwell B , Hutchison D C , Baum H . Biochem J 1987; 243: 219-223. 46. Claesson R, Karlsson M, Zhang Y , Carlsson J. J Leukoc Biol 1996; 60: 598-602. 47. Okamoto T, Akaike T, Nagano T, Miyajima S, Suga M , Ando M, Ichimori K , Maeda H . Arch Biochem Biophys 1997; 342: 261-274. 48. Odekon L E , Frewin M B , Del Vecchio P, Saba T M , Gudewicz PW. Immunology 1991; 74: 114-120. 49. Riley DJ, Berg R A , Soltys R A , Kerr JS, Guss H N , Curran SF, Laskin D L . Exp Lung Res 1988; 14: 549-563. 50. Laskin D L , Kimura T, Sakakibara S, Riley DJ, Berg R A . J Leukoc Biol 1986; 39: 255-266. 51. Parnham M J . Biochem Pharmacol 1999; 58: 209-215. 52. Pou S, Gunther M R , Pou WS, Cao G L , Bator J M , Cohen M S , Burch R M , Rosen G M . Biochem Pharmacol 1993; 45: 2123-2127. 53. González-Alvaro I, Carmona L, Díaz-González F, González-Amaro R, Mollinedo F, Sánchez-Madrid F, Laffón A, García-Vicuña R. J Rheumatol 1996; 23: 723-729. 54. Bouma M G , Jeunhomme T M , Boyle D L , Dentener M A , Voitenok N N , van den Wildenberg F A , Buurman W A . J Immunol 1997; 158: 5400-5408. 55. Díaz-González F, González-Alvaro I, Campanero M R , Mollinedo F, del Pozo M A , Muñoz C , Pivel JP, Sánchez-Madrid F. J Clin Invest 1995; 95: 1756-1765. 56. de Mello SB, Laurindo IM, Cossermelli W. Rev Paul Med 1994; 112: 489-494. 57. Blackwood R A , Hessler RJ. J Leukoc Biol 1995; 58: 114-118. 58. Currie M S , Rao K M , Padmanabhan J, Jones A, Crawford J, Cohen HJ. J Leukoc Biol 1990; 47: 244-250. 59. Ramamurthy NS, Kucine A J , McClain SA, McNamara T F , Golub LM. Adv Dent Res 1998; 12: 144-148. 60. Matheson NR, van Halbeek H , Travis J. J Biol Chem 1991; 266: 13489-13491. 61. Borregaard N, Jensen HS, Bjerrum OW. Agents Actions 1987; 22: 255-260. 62. Le Cabec V , Maridonneau-Parini I. J Biol Chem 1995; 270: 2067-2073. 63. Miesel R, Kurpisz M, Kroger H . Free Radie Biol Med 1996; 20: 75-81.

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.

75

64. Sorsa Τ, Ding Y , Salo Τ, Lauhio A , Teronen O, Ingman T, Ohtani H , Andoh N , Takeha S, Konttinen Y T . Ann Ν Y Acad Sci 1994; 732: 112-131.

Downloaded by KTH ROYAL INST OF TECHNOLOGY on March 4, 2016 | http://pubs.acs.org Publication Date: August 10, 2001 | doi: 10.1021/bk-2001-0792.ch005

65. Brandstetter H , Engh R A , Von Roedern E G , Moroder L , Huber R, Bode W, Grams F. Protein Sci 1998; 7: 1303-1309. 66. Vender R L . J Investig Med 1996; 44: 531-539.

Edwards and Vigo; Bioactive Fibers and Polymers ACS Symposium Series; American Chemical Society: Washington, DC, 2001.